Structured Biodegradable Polymeric Microparticles for Drug Delivery

Oct 1, 2015 - School of Energy, Environment & Agrifood (SEEA), Department of Offshore, Process & Energy Engineering, Cranfield University, Cranfield, ...
0 downloads 21 Views 6MB Size
Subscriber access provided by UNIV LAVAL

Article

Structured biodegradable polymeric microparticles for drug delivery produced using flow focusing glass microfluidic devices Ekanem E Ekanem, Seyed Ali Nabavi, Goran T. Vladisavljevic, and Sai Gu ACS Appl. Mater. Interfaces, Just Accepted Manuscript • DOI: 10.1021/acsami.5b06943 • Publication Date (Web): 01 Oct 2015 Downloaded from http://pubs.acs.org on October 2, 2015

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

ACS Applied Materials & Interfaces is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Structured biodegradable polymeric microparticles for drug delivery produced using flow focusing glass microfluidic devices Ekanem E. Ekanem, † Seyed Ali Nabavi, ‡Goran T. Vladisavljević,*, † and Sai Gu # †

Department of Chemical Engineering, Loughborough University, Loughborough, LE11

3TU, United Kingdom. ‡

School of Energy, Environment & Agrifood (SEEA), Department of Offshore, Process &

Energy Engineering, Cranfield University, Cranfield, MK43 0AL, United Kingdom. #

Department of Chemical and Process Engineering, Faculty of Engineering and Physical

Sciences, University of Surrey, Guildford, GU2 7XH, United Kingdom. *

Corresponding author’s address: Department of Chemical Engineering, Loughborough

University, Loughborough, LE11 3TU, United Kingdom. Phone number +441509222518; fax number +441509223923; email: [email protected]

ABSTRACT: Biodegradable poly(DL-lactic acid) (PLA) and poly(lactic-co-glycolic acid) (PLGA) microparticles with tunable size, shape, internal structure and surface morphology were produced by counter-current flow focusing in axisymmetric (3D) glass capillary devices. The dispersed phase was composed of 0.5-2 wt% polymer solution in a volatile

ACS Paragon Plus Environment

1

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 40

organic solvent (ethyl acetate or dichloromethane) and the continuous phase was 5 wt% aqueous poly(vinyl alcohol) solution. The droplets with a coefficient of variation in dripping regime below 2.5 % were evaporated to form polymeric particles with uniform sizes ranging between 4-30 µm. The particle microstructure and surface roughness were modified by adding nanofiller (montmorillonite nanoclay) or porogen (2-methylpentane) in the dispersed phase to form less porous polymer matrix or porous particles with golf-ball-like dimpled surface, respectively. The presence of 2-4 wt% nanoclay in the host polymer significantly reduced the release rate of paracetamol and prevented the early burst release, as a result of reduced polymer porosity and tortuous path for the diffusing drug molecules. Numerical modelling results using the volume of fluid-continuum surface force model agreed well with experimental behaviour and revealed trapping of nanoclay particles in the dispersed phase upstream of the orifice at low dispersed phase flow rates and for 4 wt% nanoclay content, due to vortex formation. Janus PLA/PCL (polycaprolactone) particles were produced by solvent evaporation-induced phase separation within organic phase droplets containing 3 % (v/v) PLA/PCL (30/70 or 70/30) mixture in dichloromethane. A strong preferential adsorption of Rhodamine 6G dye onto PLA was utilized to identify PLA portions of the Janus particles by Confocal Laser Scanning Microscopy (CLSM). Uniform hemispherical PCL particles were produced by dissolution of PLA domes with acetone. KEYWORDS: Microfluidic flow focusing, Biodegradable microspheres, Drug delivery systems, Poly(lactic acid), Poly(lactic-co-glycolic acid), Nanoclay, Janus particle, Hemispherical particle.

ACS Paragon Plus Environment

2

Page 3 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

INTRODUCTION Conventional oral and intravenous drug administration routes are characterized by rapid drug release and absorption.1 Limitations such as poor patient compliance (due to missing or altering dosages) and difficulty in attainment of steady state conditions (as a result of peakvalley plasma concentration fluctuations) have led to poor drug efficacy and toxicity as consequences of underdosing and overdosing of drugs respectively in patients.2 Drug delivery systems have been designed to overcome these limitations and extend, delay and target drug release3–6. When properly designed, controlled drug delivery systems should be able to deliver drugs at a predetermined rate and manner, either locally or systemically, for a specific period of time after drug administration, which can eliminate excessive fluctuations of drug plasma concentrations.7 Biocompatible polymeric materials such as non-biodegradable hydrophobic polymers,8 hydrogels,9 water soluble polymers,10 and synthetic biodegradable polymers have been widely used as vehicles for drug delivery.1 Biodegradable polymers have been mainly used in the form of microspheres, since they offer high surface area for adhesion and drug release and a low drag force during mobility in fluids.11 The most common biodegradable synthetic polymers are poly(lactic acid) (PLA), poly(lactic-co-glycolic acid) (PLGA)12, and polycaprolactone (PCL),1,13 due to their approval by FDA (Food and Drug Administration)14 and environmentally friendly nature. In order to meet different dosage requirements, there is a strong need to modify drug solubility and release rate by incorporation of different additives into a host biodegradable polymer.

ACS Paragon Plus Environment

3

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 40

Nanoclays are FDA approved plate-like nanoparticles of layered mineral silicates that can be integrated into the polymer matrix to improve mechanical and rheological properties of polymeric microspheres and increase the solubility and bioavailability of drugs such as ibuprofen, indomethacin, diclofenac, naproxen, fenbufen, and paracetamol.2 Nanoclays also have highly desirable dietary, bacterial and metabolic detoxification characteristics15. Another method of modification of drug release kinetics is internal phase separation caused by addition of non-solvent in the mixture of polymer and solvent prior to emulsification. Depending on polymer to non-solvent ratio and the type of solvent, non-solvent and polymer, phase separation within emulsion droplets (triggered by solvent evaporation) can lead to core/shell, occluded, acorn or heteroaggregated particle morphology.16 Hole-shell and crescent-moon-shaped microparticles can be synthesised from acorn-shaped biphasic droplets by polymerisation of one of the phases.17,18 Such unique morphologies can modify drug release profile12,19,20 and increase the potential for alternative routes of drug administration. Emulsion-solvent evaporation is one of the most common methods for fabrication of polymeric microspheres from pre-formed polymers. Emulsification can be carried out using conventional top-down methods based on breaking larger droplets into smaller ones,21 such as mechanical agitation, high-pressure homogenization, and sonication or bottom-up methods, based on direct drop-by-drop generation, such as membrane emulsification22–24, microchannel emulsification,25 and microfluidic techniques.26–29 Conventional emulsification methods are energy intensive and often lead to polydispersed microspheres. Microfluidic methods allow production of particles with low polydispersity (CV = Standard Deviation/Mean < 3%) and high encapsulation efficiency that are perfectly tailored to meet the needs of pharmaceutical industry.30 However, microfluidic droplet generators are typically planar (2-D) and made

ACS Paragon Plus Environment

4

Page 5 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

from single crystal silicon chips using expensive microfabrication methods31 or mouldable polymers such as polydimethylsiloxane (PDMS), which swell and deform in contact with organic solvents.29 Here, we investigate formation of structured polymeric microspheres using glass capillary microfluidic devices. These devices pioneered by Utada et al.32 offer 3-D flow focusing that minimizes wetting, they are cheap to fabricate and more mechanically robust and inert than polymeric devices.33,13 Microstructural modifications of polymer matrix occurred due to introduction of nanoclay, non-solvent or second polymer, which resulted in the formation of composite and non-spherical particles and particles with structured surface (Figure 1b-d).

Figure 1. Strategies used in this work for creation of polymer particles: (a) Plain polymer particle; (b) Nanoclay embedded polymer particle; (c) Golf ball-like particle; (d) Janus and hemispherical particle.

ACS Paragon Plus Environment

5

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 40

A special attention was paid on the effect of flow conditions on the retention (trapping) of suspended nanoparticles in the device. This phenomenon is important in all microfluidic processes where fluid streams entering a microfluidic channel contain suspended nanoobjects, such as colloidal particles, nano-assemblies, nanofibers, nanotubes or viruses. The effect of internal structure of the particles on the release rate of encapsulated drug was also discussed. Glass capillary devices have been used for fabrication of vesicles such as colloidosomes34, polymersomes,35 and liposomes,36 and microparticles such as solid lipid37 and microgel33 particles. This paper presents new applications of glass capillary microfluidics for the production of drug delivery vehicles composed of nanoclay-embedded, golf ball-like, Janus, and hemispherical polymeric particles. A facile control over the particle size and morphology achieved in this work enables accurate prediction of drug release behavior.

EXPERIMENTAL SECTION Materials. Poly(dl-lactic acid) (PLA, IngeoTM 4060D, Mw = 89,000 gmol-1) supplied by NatureWorks LLC (Minnetonka, MN, USA), poly(lactic-co-glycolic acid) (PLGA, Mw = 10,000 g mol-1, 50% DL-lactic acid and 50% glycolic acid), supplied by Wako Pure Chemical Industries (Osaka, Japan) and polycaprolactone (PCL, Mw = 14,000 g mol−1) supplied by Sigma-Aldrich (UK) served as primary ingredients of the microspheres. Organically modified nanoclay (Cloisite ® 30B, ρ = 1980 kg m-3, Southern Clay Products, Gonzales, TX, USA) consisting of plate-like montmorillonite nanoparticles surface-substituted with alkyl

ACS Paragon Plus Environment

6

Page 7 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

ammonium ions, was used as a polymer filler for the microspheres. Dichloromethane (DCM, HPLC grade, Fisher Scientific, UK) and ethyl acetate (Sigma-Aldrich, UK) were used as nonpolar polymer solvents. 2-Methylpentane (Sigma-Aldrich, UK) was added to the dispersed phase as a polymer non-solvent in the production of golf ball-like particles. Paracetamol (4acetamidophenol) obtained from Sigma-Aldrich was used for drug release experiments. Poly(vinyl alcohol) (PVA, Mw = 13,000-23,000, 87-89% hydrolyzed) supplied by SigmaAldrich (UK) was used as a water-soluble stabilizer in the continuous phase. Reverse osmosis water was prepared using Millipore 185 Milli-Q Plus apparatus. Rhodamine 6G (Rh6G) and Nile red (Sigma-Aldrich, UK) were used as fluorescence dyes for identification of bifacial particle

morphology.

2-[methoxy(polyethylenoxy)propyl]-trimethoxysilane

(MPEOPS)

supplied by Fluorochem Ltd (UK) was used to enhance hydrophilicity of inner capillary. Methods. Fabrication of Glass Capillary Device. A round capillary with inner diameter 0.58 mm and outer diameter 1 mm (Intracel, UK) was pulled using a micropipette puller (Sutter Instrument Company, USA). The tip of the pulled capillary was polished to desired orifice size using an abrasive paper (Black Ice Waterproof T402 Paper, Alpine Abrasives, UK), cleaned with compressed air and Milli-Q water and treated with MPEOPS. A 150-mm long square capillary with an inner diameter of 1.05±0.1 mm (Atlantic International Technologies, USA) was cut with a pen cutter to a required length of 50 mm and attached to a microscopic slide using epoxy glue (5-Minute Epoxy® Devcon). The round capillary was then inserted in the square capillary and centred using an inverted microscope (XDS-3, GX Microscopes, UK) before it was glued to position. The exposed end of the round capillary was connected via PVC medical tubing (ID: 1.00 mm, OD: 1.05 mm; Fisher Scientific, UK) to collection

ACS Paragon Plus Environment

7

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 40

vial. All fabrication steps were carried out in a dust-free environment and disposable medical gloves were worn to prevent contamination of the device. Production of emulsions and particles. The compositions of dispersed phase used in the production of different particles are shown in Table 1. Table 1. Composition of dispersed phase used in this work for preparation of different particles. The continuous phase was 5 wt% PVA in Milli-Q water in all formulations. Particle structure Plain polymer particles Nanoclay-embedded polymer particles Golf ball-like polymer particles Janus polymer particles

Dispersed phase composition

wt%

Polymer (PLA or PLGA) Solvent (DCM or ethyl acetate) Nanoclay Polymer (PLA or PLGA) Solvent (DCM or ethyl acetate) Nonsolvent (2-methylpentane) Polymer (PLA or PLGA) Solvent (DCM or ethyl acetate) Polymer mixture (PCL + PLA) (PCL:PLA = 1:2 or 2:1 v/v) Solvent (DCM)

0.5, 1 or 2 99.5, 99 or 98 0.02 or 0.04 0.98 or 0.96 99 0.3 or 3 0.7 or 7 99 or 90 3 97

ACS Paragon Plus Environment

8

Page 9 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

Figure 2. (a) Experimental setup of microfluidic device for the production and monitoring of microdroplet generation; (b) and (c) Monodispersed droplet generation in the dripping regime for a dispersed phase of PLA/nanoclay in DCM; (d) Polydispersed droplet generation in the jetting regime for dispersed phase of PLA in DCM. Scale bars: 250 µm.

For droplet production, glass capillary device was placed on a GXM XD63 optical inverted microscope. The dispersed and continuous phase of desired formulation (Table 1) were delivered from two gas tight syringes with Luer-lock fitting (VWR Catalyst Company, UK) mounted on separate Harvard Apparatus 11 Elite syringe pumps (Figure 2). PTFE medical tubing (Smiths Medical International Ltd, UK) and polyethylene medical tubing (Fisher Scientific, UK) were used to deliver the dispersed and continuous phase respectively to the corresponding needle on the device. The continuous phase was delivered via a 3-way stop cock (VWR Catalyst Company, UK) and Puradisc syringe filter (Sigma-Aldrich Company, UK) for easy syringe refilling and removal of microscopic impurities in the liquid. The two phases were supplied within the device counter-currently at low flow rates, such that the oil/aqueous interface slowly approached the orifice of the inner capillary to avoid wetting the tip by the oil phase. Droplet formation occurred on adjustment flowrates with syringe pumps. Droplets were generated within a tapered section of the collection capillary (Figure 2 b-d). Real-time monitoring of droplet generation was achieved with a high speed camera (Phantom V9.0, Vision Research, USA) supported by the Phantom Camera Control (PCC) software (version 2.14.727.1, Vision Research, USA), which provided interface for image acquisition, brightness/contrast adjustment, and post-generation analysis. Generated droplets were collected in a vial and vacuum evaporated at room temperature to form particles.

ACS Paragon Plus Environment

9

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 40

Analysis of Video Recordings. Recorded videos of droplet generation were stored in the camera’s internal memory with the aid of the PCC software. The number of frames per second  and the number of frames stored  were chosen based on the storage space

availability and droplet generation rate. The droplet generation frequency  was calculated as:

 =   / , where  is the number of droplets generated in  / seconds, determined using ImageJ v.1.44 software (Wayne Rasband, National Institute of Health).

X-Ray Diffraction (XRD) Analysis. Wide angle X-ray diffraction patterns were obtained using a Bruker D2 Phaser diffractometer fitted with a 1-dimensional LynxEYETM detector. A copper X-ray source (Kα = 1.54184 Å) was used run at 30 kV and 10 mA, with Kβ radiation suppressed by means of a 0.5 mm thick nickel filter. Patterns were recorded over a 2θ range of 2-40o with a step size of 0.02o and an equivalent step time of 49.2 s. Sample rotation was set at 15 rpm. Bruker’s proprietary Eva 2.0 software was used to obtain the spectra. Scanning Electron Microscopy (SEM) and Focused Ion Beam (FIB) Imaging. Imaging was carried out using a dual-beam focused ion beam scanning electron microscope (FIB-SEM) (Nova 600 Nanolab, FEI Company, Hillsboro, Oregon, USA). All SEM imaging were carried out at 10 kV and 2.1 nA over a 30 µ aperture exposure. For FIB imaging 30 kV was used throughout. External angular imaging was initially obtained at 30 pA before an increase to 0.3 nA for a protective platinum (Pt) deposit. A cross-section was then obtained at 20 nA before cross section cleaning was carried out at 7 nA and reduced to 3nA for a final cleaning. Cleaning was necessary for better microstructural imaging. The current was finally reduced to 30 pA to preserve the exposed cross-sectional microstructure

ACS Paragon Plus Environment

10

Page 11 of 40

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Materials & Interfaces

and prevent modification due to beam exposure during imaging, which was carried out at approximately 2 min/image for noise reduction. Confocal Laser Scanning Microscopy (CLSM). The microstructure of Janus particles was vizualised using a Nikon Eclipse TE300 confocal inverted microscope connected to a computer running Zeiss LaserSharp 2000™ software. The Janus particles were stained with Rhodamine 6G and Nile Red, added to the dispersed phase prior to emulsification. A suspension of Janus particles to be scanned was placed on a microscopic slide and allowed to dry. The sample was excited with argon laser with a wavelength of 488 nm and helium-neon laser with a wavelength of 543 nm. The total emission was separated into two images captured by two photomultiplier tubes (PMTs): PMT1 captured fluorescence between 515±30 nm (green region) and PMT2 captured fluorescence above 570 nm (yellow-red region). Cumulative Drug Release Study. 0.7 wt% of paracetamol (PCM) was added to the dispersed phase containing ethyl acetate (EA). The concentration of PCM in the continuous phase was measured using a Lambda 35 UV/Vis (PerkinElmer Instruments, UK) double beam spectrophotometer, supported by UVWINLAB version 2.85.04 software (PerkinElmer Instruments, UK), over 120 h at a wavelength of 242 nm. The solubility of PCM in EA and water is 10.7 and 17.4 g/kg, respectively.38 The cumulative release was determined by:     =

 × 100 1  

 where  is the concentration of drug in the continuous phase and  is the predicted

drug concentration in the continuous phase after complete release.

ACS Paragon Plus Environment

11

ACS Applied Materials & Interfaces

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 40

Governing Equations and Numerical Modelling. Numerical modelling of drop generation was carried out using a two-dimensional axisymmetric volume of fluid (VOF) approach.39 The following equations of mass, Eq. 2, and momentum conservation, Eq. 3, were used:  %&' = 0 2 + ! ∙ #$ 

 %&' + ! ∙ #$ %&$ %&' = −!* + ! ∙ +,#!$ %& + !$ %& - '. + / %%%%&0 3 #$ 

%& and P are velocity and pressure and t, ,, and  are time, dynamic viscosity and where $

density. The interfacial force, /2 , and gravitational force are included in the source term, /0 , However, since the length-scale is in the order of micro, the gravitational force is negligible %%%&2 . In VOF model, a momentum equation is solved for all phases and the advection /0 = / and %%%%&

of interface is tracked by solving a transport equation for volume fraction, :

 %&. ! = 0 4 +$ 

where f determines the portion of each cell filled with one of the two phases: f=0

the cell is filled with aqueous phase

0